Underwater acoustic modeling: principles, techniques and applications [2 ed.] 0419201904, 9780419201908

The subject of underwater acoustic modeling deals with the translation of our physical understanding of sound in the sea

260 21 8MB

English Pages 303 Year 1995

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Underwater acoustic modeling: principles, techniques and applications [2 ed.]
 0419201904, 9780419201908

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Underwater Acoustic Modeling

Underwater Acoustic Modeling Principles, techniques and applications Second edition

Paul C.Etter

E & FN SPON An Imprint of Chapman & Hall London · Glasgow · Weinheim · New York · Tokyo · Melbourne · Madras

Published by E & FN Spon, an imprint of Chapman & Hall, 2±6 Boundary Row, London SE1 8HN, UK Chapman & Hall, 2–6 Boundary Row, London SE1 8HN, UK Blackie Academic & Professional, Wester Cleddens Road, Bishopbriggs, Glasgow G64 2NZ, UK Chapman & Hall GmbH, Pappelallee 3, 69469 Weinheim, Germany Chapman & Hall USA, 115 Fifth Avenue, New York, NY 10003, USA Chapman & Hall Japan, ITP-Japan, Kyowa Building, 3F, 2–2–1 Hirakawacho, Chiyoda-ku, Tokyo 102, Japan Chapman & Hall Australia, 102 Dodds Street, South Melbourne, Victoria 3205, Australia Chapman & Hall India, R.Seshadri, 32 Second Main Road, CIT East, Madras 600 035, India First edition 1991 This edition published in the Taylor & Francis e-Library, 2005. “To purchase your own copy of this or any of Taylor & Francis or Routledge’s collection of thousands of eBooks please go to www.eBookstore.tandf.co.uk.” © Elsevier Science Publishers Ltd. Second edition 1996 © 1996 Paul C.Etter ISBN 0-203-47565-8 Master e-book ISBN

ISBN 0-203-78389-1 (Adobe eReader Format) ISBN 0 419 20190 4 (Print Edition) Apart from any fair dealing for the purposes of research or private study, or criticism or review, as permitted under the UK Copyright Designs and Patents Act, 1988, this publication may not be reproduced, stored, or transmitted, in any form or by any means, without the prior permission in writing of the publishers, or in the case of reprographic reproduction only in accordance with the terms of the licences issued by the Copyright Licensing Agency in the UK, or in accordance with the terms of licences issued by the appropriate Reproduction Rights Organization outside the UK. Enquiries concerning reproduction outside the terms stated here should be sent to the publishers at the London address printed on this page. The publisher makes no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal responsibility or liability for any errors or omissions that may be made. A catalogue record for this book is available from the British Library

To my wife Alice and my sons Gregory and Andrew

Contents

Preface

ix

Preface to the first edition

xi

Acknowledgments

xiii

1

Introduction

1

1.1

Background

1

1.2

Measurements and prediction

2

1.3

Developments in modeling

4

1.4

Inverse acoustic sensing of the oceans

5

Acoustical oceanography

8

2.1

Background

8

2.2

Physical and chemical properties

8

2.3

Sound speed

11

2.4

Boundaries

16

2.5

Dynamic features

27

2.6

Biologics

39

Propagation I. Observations and physical models

41

3.1

Background

41

3.2

Nature of measurements

42

3.3

Basic concepts

43

3.4

Sea-surface boundary

47

3.5

Sea-floor boundary

52

3.6

Attenuation and absorption

58

3.7

Surface ducts

58

3.8

Deep sound channel

64

2

3

vi

3.9

Convergence zones

65

3.10

Reliable acoustic path

67

3.11

Shallow-water ducts

69

3.12

Arctic half-channel

70

3.13

Coherence

72

Propagation II. Mathematical models (Part One)

74

4.1

Background

74

4.2

Theoretical basis for propagation modeling

75

4.3

Ray theory models

77

4.4

Normal mode models

86

4.5

Multipath expansion models

91

4.6

Fast-field models

92

4.7

Parabolic equation models

93

4.8

The RAYMODE model—a specific example

98

4.9

Numerical model summaries

102

Propagation II. Mathematical models (Part Two)

109

5.1

Background

109

5.2

Surface duct models

109

5.3

Shallow-water duct models

114

5.4

Arctic models

124

5.5

Data support requirements

126

5.6

Special applications

131

Noise I. Observations and physical models

143

6.1

Background

143

6.2

Noise sources and spectra

143

6.3

Depth dependence

149

6.4

Directionality

150

6.5

Arctic ambient noise

151

6.6

Acoustic daylight

152

Noise II. Mathematical models

154

4

5

6

7

vii

7.1

Background

154

7.2

Theoretical basis for noise modeling

154

7.3

Ambient noise models

155

7.4

The RANDI model—a specific example

157

7.5

The noise notch

159

7.6

Beam noise statistics models

163

7.7

Data support requirements

164

7.8

Numerical model summaries

164

Reverberation I. Observations and physical models

168

8.1

Background

168

8.2

Volume reverberation

169

8.3

Boundary reverberation

171

Reverberation II. Mathematical models

178

9.1

Background

178

9.2

Theoretical basis for reverberation modeling

178

9.3

Cell-scattering models

181

9.4

The REVMOD model—a specific example

183

9.5

Bistatic reverberation

187

9.6

Point-scattering models

190

9.7

Numerical model summaries

193

10

Sonar performance models

194

10.1

Background

194

10.2

Sonar equations

195

10.3

The NISSM model—a specific example

197

10.4

Model-operating systems

205

10.5

Data sources and availability

212

10.6

Numerical model summaries

217

Model evaluation

218

11.1

Background

218

11.2

Past evaluation efforts

219

8

9

11

viii

11.3

Analytical benchmark solutions

220

11.4

Quantitative accuracy assessments

222

11.5

The POSSM experience—a specific example

223

11.6

Evaluation guidelines

226

11.7

Documentation standards

229

Appen dix A:

Abbreviations and acronyms

232

Appen dix B:

Glossary of terms

241

References

247

Author index

274

Subject index

282

Preface

The subject of underwater acoustic modeling deals with the translation of our physical understanding of sound in the sea into mathematical formulas solvable by computers. This book divides the subject of underwater acoustic modeling into three fundamental aspects: the physical principles used to formulate underwater acoustic models; the mathematical techniques used to translate these principles into computer models; and modeling applications in sonar technology and oceanographic research. The material presented here emphasizes aspects of the ocean as an acoustic medium. It shows mathematicians and physical scientists how to use this information to model the behavior of sound in a spatially complex and temporally variable ocean. This approach diminishes the need for discussions of engineering issues such as transducers, arrays and targets. Aspects of hardware design and modeling in underwater acoustics are discussed in other excellent texts. Recent developments in underwater acoustic modeling have been influenced by changes in global geopolitics. These changes are evidenced by strategic shifts in military priorities as well as by efforts to transfer defense technologies to non-defense applications. The strategic shift in emphasis from deep-water to shallow-water naval operations has focused attention on improving sonar performance in coastal regions. These near-shore regions, which are sometimes referred to as the littoral zone, are characterized by complicated and highly variable acoustic environments. Such difficult environments challenge the abilities of those sonar models intended for use in deep-water scenarios. This situation has prompted further development of underwater acoustic models suitable for forecasting and analyzing sonar performance in shallow-water areas. The policy of defense conversion has encouraged the transfer of sonar modeling technology to nondefense applications. Much of this transfer has benefited the growing field of environmental acoustics, which seeks to expand exploration of the oceans through acoustic sensing. Such technology conversion is exemplified by the utilization of naval underwater acousic models as both prognostic and diagnostic tools in sophisticated experiments employing inverse acoustic sensing of the oceans. These rapid developments in modeling have created a need for a second edition. The intent is to update recent advances in underwater acoustic modeling and to emphasize new applications in oceanographic research. This edition also reflects a broader international interest in the development and application of underwater acoustic models. The coming years promise to be challenging in terms of defining research directions, whether for defense or industry, and this edition should provide technology planners with a useful baseline. The original organization of material into eleven chapters has served well and therefore remains unchanged. When required, new material has been arranged into additional subsections. Comments from users of the first edition have evidenced appeal from acousticians, as well as oceanographers, who have enthusiastically endorsed this book as both a practical tool and an instructional

x

aid. In this latter regard, several academic institutions have utilized this book as an adjunct text for graduatelevel courses in applied mathematics and ocean sciences. This edition has benefited from a continuation of my short courses which, since 1993, have been offered through the Applied Technology Institute of Clarksville, Maryland (USA). Continued exposure to the insightful questions posed by my students has provided me with the opportunity to further refine my presentation. Despite the appearance of several new books in the field of ocean acoustics, this book remains unique in its treatment and coverage of underwater acoustic modeling. It is a pleasure to note that the first edition has been recognized as an authoritative compendium of state-of-the-art models and is often cited as the standard reference. Paul C.Etter Rockville, Maryland USA

Preface to the first edition

The subject of underwater acoustic modeling deals with the translation of our physical understanding of sound in the sea into mathematical formulas solvable by computers. These models are useful in a variety of research and operational applications including undersea defense and marine seismology. There has been a phenomenal growth in both the number and types of models developed over the past several decades. This growth reflects the widespread use of models for the solution of practical problems as well as the considerable advances made in our computational abilities. The primary motivation for the development of underwater acoustic models is defense related. Researchers involved in Anti-Submarine Warfare (ASW) and associated undersea defense disciplines use models to interpret and forecast acoustic conditions in the sea in support of sonar design and sonar operation. Consequently, the emphasis in this book is placed on those models that are particularly useful in solving sonar performance problems. Users and potential users of models are commonly ill acquainted with model formulations. As a result, the capabilities and limitations of the models are poorly understood and the models are often improperly used. Moreover, the sheer number of available models complicates the process of model selection. This book is intended for those who have a fundamental understanding of underwater acoustics but who are not familiar with the various aspects of modeling. Sufficient mathematical derivations are included to demonstrate model formulations, and guidelines are provided to assist in the selection and proper application of these models. Comprehensive summaries identify the available models and associated documentation. The material in this book is organized into eleven chapters. The order of presentation follows the structure suggested by a hierarchical method of sonar model construction. Chapter 1 introduces the types of underwater acoustic models, provides a framework for the consistent classification of modeling techniques, and defines the terminology common to modeling work. Aspects of oceanography essential to an understanding of acoustic phenomena are presented in Chapter 2. Chapters 3 through 9 address the observations and models dealing with propagation, noise and reverberation in the sea. In Chapter 10, the information from Chapters 3–9 is integrated into sonar performance models. Finally, Chapter 11 describes the process of model evaluation. Since 1982, I have developed and taught a series of intensive short courses for the Technology Service Corporation of Silver Spring, Maryland (USA). Earlier versions of this course were taught in collaboration with Professor Robert J.Urick of the Catholic University of America. Professor Urick would discuss underwater acoustic measurements while I would review the related modeling techniques. As the course evolved into one in which I became the sole instructor, I borrowed heavily from Professor Urick’s several books (with permission) in order to preserve the continuity of the course material. The success of this course encouraged me to publish my class notes as a book.

xii

Many notable books have been published in the field of underwater acoustics. None, however, has dealt exclusively with modern developments in modeling, although some have addressed aspects of propagation modeling. This book is unique in that it treats the entire spectrum of underwater acoustic models including environmental, propagation, noise, reverberation and sonar performance models. I have intentionally preserved the notation, terminology and formalism used by those researchers whose work I have cited. I have also intentionally emphasized aspects of oceanography since my experience has indicated that many acousticians have little appreciation for the complex role played by the ocean as an acoustic medium. Conversely, oceanographers frequently fail to appreciate the great potential of underwater acoustics as a remote sensing technique. Paul C.Etter

Acknowledgments

The students who have attended my short courses have provided both a receptive and critical audience for much of the material contained in this book. Many of my colleagues have provided suggestions and useful insights: Dr Aubrey L.Anderson, Dr Stanley A.Chin-Bing, Dr Richard B. Evans, Dr Robert W.Farwell, Dr Richard P.Flanagan, Dr Robert L.Martin, Dr Frederick D.Tappert and Dr Henry Weinberg. Robert S.Winokur provided administrative guidance in the early stages of my work in underwater acoustic modeling. Professor Robert J.Urick has been a source of great encouragement and has graciously allowed me to liberally borrow material from his several books. Professor John D.Cochrane of Texas A & M University inspired the scholarly discipline that facilitated creation of this book.

1 Introduction

1.1 BACKGROUND The field of underwater acoustics has been extensively developed over the past four decades in response to practical needs originating within both the naval sonar and the marine seismology communities. The emphasis in this book is focused on those underwater acoustic models developed for sonar (versus seismic) applications. Since 1960 considerable work has been expended on the development of models with which to analyze data collected in field experiments. These models have then been used to forecast acoustic conditions for planning at-sea experiments, designing optimized sonar system and predicting sonar performance at sea. Modeling has become the chief mechanism by which researchers and analysts can simulate sonar performance in laboratory conditions. Modeling provides an efficient means by which to parametrically investigate the performance of hypothetical sonar designs under varied environmental conditions as well as to estimate the performance of existing sonars in different ocean areas and seasons. A distinction is made between physical models and mathematical models, both of which are addressed in this book. Physical models pertain to theoretical or conceptual representations of the physical processes occurring within the ocean; the term ‘analytical model’ is sometimes used synonymously. Mathematical models include both empirical models (those based on observations) and numerical models (those based on mathematical representations of the governing physics). The subject of analog modeling, which is defined here as controlled acoustic experimentation in water tanks employing appropriate oceanic scaling factors, is not addressed in this book. Reviews of acoustic analog modeling have been presented by Barkhatov (1968) and Zornig (1979). The physical models underlying the numerical models have been well known for some time. The transition to operational computer models, however, has been hampered by several factors: limitations in computer capabilities; inefficient mathematical methods; and inadequate oceanographic and acoustic data with which to initialize and evaluate models. This book addresses three broad types of underwater acoustic models: environmental models, basic acoustic models and sonar performance models. The first category, environmental models, includes empirical algorithms which are used to quantify the boundary conditions (surface and bottom) and volumetric effects of the ocean environment. Such models include sound speed, absorption coefficients, surface and bottom reflection losses, and surface, bottom and volume backscattering strengths.

2

INTRODUCTION

The second category, basic acoustic models, comprises propagation (transmission loss), noise and reverberation models. This category is the primary focus of attention in this book. The third category, sonar performance models, is composed of environmental models, basic acoustic models and appropriate signal-processing models. Sonar performance models are organized to solve specific sonar applications problems such as submarine detection, mine hunting, torpedo homing and bathymetric sounding. Figure 1.1 illustrates the relationships among these three broad categories of models. As the applications become more and more system specific, that is, as one progresses from environmental models toward sonar performance models, the respective models become less universal in application. This is a consequence of the fact that system-specific characteristics embedded in the higher level models (for example, signalprocessing models) restrict their utility to particular sonar systems. Thus, while one propagation model may enjoy a wide variety of applications, any particular sonar performance models is, by design, limited to a relatively small class of well-defined sonar problems. The wide breadth of material covered in this book precludes exhaustive discussions of all existing underwater acoustic models. Accordingly, only selected models considered to be representative of each of the three broad categories will be explored in more detail. In addition, comprehensive summaries identifying existing basic acoustic models and sonar performance models are presented. Notable environmental models are identified and discussed in appropriate sections throughout this book. Modeling applications will generally fall into one of two basic areas: research or operational. Researchoriented applications are conducted in laboratory environments where accuracy is important and computer time is not a critical factor. Examples of research applications include sonar system design and field experiment planning. Operationally oriented applications are conducted as field activities, including fleet operations at sea and sonar system training ashore. Operational applications generally require rapid execution, often under demanding conditions; moreover, modeling accuracy may be subordinate to processing speed. 1.2 MEASUREMENTS AND PREDICTION The scientific discipline of underwater acoustics is transitioning from a stage of observation to a stage of understanding and prediction. This transition is not always smooth—direct observations are limited, the resulting prediction tools (models) are not perfected, and much refinement is still needed. Experimental measurements in the physical sciences are generally expensive due to instrumentation and facility operation costs. In the case of oceanographic and underwater acoustic data collection, this is particularly true because of the high costs of platform operation (ships, aircraft, submarines). Consequently, in the field of underwater acoustics, much use is made of what few field measurements already exist. Notable large-scale field programs that have been successfully conducted in the past include AMOS (Acoustic, Meterological and Oceanographic Survey) and LRAPP (Long Range Acoustic Propagation Project). Modeling has been used extensively to advance scientific understanding without expending inordinate resources on additional field observations. The balance between observations and modeling, however, is very delicate. Experimenters agree that modeling may help to build intuition or refine calculations, but they argue further that only field observations give rise to genuine discovery. Accordingly, many researchers find mathematical models most useful once the available observations have been analyzed on the basis of simple physical models.

MEASUREMENTS AND PREDICTION

3

Fig. 1.1 Generalized relationships among environmental models, basic acoustic models and sonar performance models.

The relationship between experimentation and modeling (in the furtherance of understanding and prediction) is depicted schematically in Fig. 1.2. Here, physical models form the basis for numerical models while experimental observations form the basis for empirical models. Moreover, analog modeling is represented as a form of laboratory (versus field) experimentation. Scientists are becoming more aware of the connection between physical processes and computation and many now find it useful to view the world in computational terms. Consequently, computer simulation is sometimes viewed as a third form of science, halfway between theory and experiment. Furthermore, understanding can be enhanced through the use of advanced computer graphics to convert large volumes of data into vivid and comprehensible patterns. Because of national security concerns, some existing data sets are limited in accessibility. Also, because of the wide range of acoustic frequencies, ocean areas and geometries of interest to researchers, it is virtually impossible to accommodate all observational requirements within normal fiscal constraints. To make matters worse, acoustic data are sometimes collected at sea without the supporting oceanographic data; thus, models cannot always replicate the observed acoustic results because they lack the necessary input parameters for initialization. Satellites, together with other remote-sensing techniques, provide a useful adjunct to the prediction of underwater acoustic conditions. Specifically, many dynamic features of the ocean affect the behavior of sound, and knowledge of their location can improve the prediction of sonar performance. Although satelliteborne sensors detect only surface (or near-surface) features of the ocean such as thermal contrast, color or surface roughness, these surface expressions can generally be associated with dynamic oceanographic features below the surface, particularly when comprehensive climatological databases already exist with which to establish such associations. Thus, for example, satellite imagery can be used to provide timely and

4

INTRODUCTION

Fig. 1.2 Schematic relationship between experimentation and modeling.

accurate position information on variable features such as ocean fronts and eddies which are known to significantly affect the propagation of acoustic signals in the sea. Tactical oceanographic data collection has been augmented by drifting buoys, which use satellite relays to transmit data to mobile or stationary receiving stations, and by Autonomous Underwater Vehicles (AUVs) to access remote ocean areas such as shallow-water and under-ice regions (Brutzman et al., 1992; Selsor, 1993; Dantzler et al., 1993). The problem of operational sonar prediction embrances many disciplines, one of which is modeling. Such modern operational applications involve not only underwater acoustic models but also oceanographic models (Etter, 1989). The coupling of these two types of models provides a valuable set of prediction tools to the force commanders by enabling them to respond to the changing environmental conditions that affect their sonar performance. The remote-sensing data now available to naval forces afloat can be used in conjunction with oceanographic models to accurately forecast the locations and characteristics of dynamic ocean features (e.g. Robinson, 1992). This information can then be input to the appropriate acoustic models to assess the resultant impacts on sonar performance. These sonar systems can then be optimized for performance in each region of operation at any given time of the year. 1.3 DEVELOPMENTS IN MODELING A goal of science is to develop the means for reliable prediction to guide decision and action (Ziman, 1978). This is accomplished by finding algorithmic compressions of observations and physical laws. Physical laws are statements about classes of phenomena, and initial conditions are state-ments about particular systems.

MEASUREMENTS AND PREDICTION

5

Thus, it is the solutions to the equations, and not the equations themselves, that provide a mathematical description of the physical phenomena. In constructing and refining mathematical theories, we rely heavily on models. At its conception, a model provides the framework for a mathematical interpretation of new phenomena. In its most elemental form, a model is intended to generalize and abstract. A perfect model is one that perfectly represents reality. In practice, however, such a perfect model would defeat its purpose: it would be as complex as the problem it is attempting to represent. Thus, modeling in the physical sciences is normally reduced to many, more easily managed, components. With the advent of digital computers, modeling in the physical sciences has advanced dramatically. Improvements in computer capabilities over the past several decades have permitted researchers to incorporate more complexity into their models, sometimes with little or no penalty in run time or computer costs (e.g. Hodges, 1987; Runyan, 1991). Frequently, models become data limited. This means that observational data are lacking in sufficient quantity or quality with which to support model initialization and model evaluation. As modeling techniques have proliferated within the underwater acoustics community, it has become increasingly difficult to take stock of the various models already in existence before launching a new effort to develop yet more models. Moreover, analysts confronted with sonar performance problems have had difficulty in determining what models existed and, of those, which were best for their particular situation. This book had its genesis in just such a dilemma. A small study was sponsored by the US Navy in 1978 (Etter and Flum, 1978) to review the availability of numerical models of underwater acoustic propagation, noise and reverberation as well as the availability of databases with which to support model development and operation. Results of this work, and extensions thereto (Etter and Flum, 1980; Etter et al., 1984), have subsequently been presented at meetings of the Acoustical Society of America (Etter and Flum, 1979; Etter, 1987b) and have been updated periodically in literature review articles (Etter, 1981, 1984, 1987a, 1990). An enhanced version of the first review article (Etter, 1981) was included as Chapter 3 in a book by Urick (1982). Collectively, this work later evolved into a series of lectures and culminated in the writing of this book. The technical literature cited in this book includes many unpublished reports (so-called ‘gray’ literature) since no other sources of technical information were available. Weston and Rowlands (1979) have reviewed the development of models with application to underwater acoustic propagation over the period 1963–1978. DiNapoli and Deavenport (1979) have provided a highly mathematical examination of a select number of propagation models. Brekhovskikh and Lysanov (1991) have presented a comprehensive Russian perspective on underwater acoustics with a limited treatment of modeling. Jensen et al. (1994) have provided a comprehensive review of recent theoretical developments in ocean acoustic propagation modeling. 1.4 INVERSE ACOUSTIC SENSING OF THE OCEANS Inverse sensing involves the extraction of information from direct measurements of physical properties. Inverse-sensing techniques that employ acoustics have been used in several subdisciplines of geophysics including seismology, meteorology and oceanography. Seismologists have used tomographic techniques to infer the bulk properties of the lith-osphere (e.g. Menke, 1989). Atmospheric scientists have employed naturally generated, low-frequency sound (microbaroms) to probe the upper layers of the atmosphere in an inverse fashion (Donn and Rind, 1971). In oceanography, inverse acoustic data provide estimates of

6

INTRODUCTION

spatially integrated, and temporally averaged, oceanic conditions that are not readily available from a traditional constellation of point sensors (e.g. Bennett, 1992). Traditionally, direct (versus inverse) acoustic sensing of the oceans has been used for sonar applications. This emphasis stems principally from the military significance of sonars in the detection of submarines. Despite the restrictiveness of military security, an extensive body of relevant research has accumulated in the open literature, and much of the recent literature addresses the development and refinement of numerical codes that model the oceans as an acoustic medium. These models are utilized either in a prognostic mode (e.g. sonar performance prediction) or in a diagnostic mode (e.g. sonar design). This situation has stimulated the formation of a new subdiscipline of acoustics known as computational ocean acoustics. Related developments have been documented by Merklinger (1987), Lee et al. (1990a, 1990b, 1990c, 1993) and Lau et al. (1993). In the commercial sector, recent applications of direct acoustic-sensing methods include: Acoustic Doppler Current Profilers (ADCPs) for measuring currents; compact sonars for obstacle location and avoidance by Autonomous Underwater Vehicles (AUVs) (e.g. Brutzman et al., 1992); fish-finding devices; underwater communications systems for divers; fathometers for bathymetric sounding and navigation; and side-scanning sonars for topographic mapping of the sea-floor relief. A recent innova-tion entails acoustic monitoring of marine environmental pollution; for example, point-source and non-point-source pollution studies now use acoustic backscatter and Doppler current meter measurements. Inverse acoustic sensing of the oceans utilizes one of three natural phenomena: propagation, noise or reverberation. These techniques and their applications are briefly summarized below. Acoustic propagation characteristics in the deep oceans are determined largely by the refractive properties of the water column and, to a lesser extent, by the surface and bottom boundary conditions. Propagation measurements can be used to infer bulk properties of the water column such as temperature, sound speed, density and currents. In shallow-ocean areas, where propagation characteristics can be strongly affected by the bottom boundary, propagation measurements can be used to infer properties of the sea floor such as composition and scattering characteristics. Inverse acoustic-sensing methods utilizing the propagation characteristics of the oceans include tomography and matched field processing (see Chapter 5). The ambient noise field in the oceans is described by the spectral, spatial and temporal characteristics of sound generated by both natural and industrial sources. Measurements of these characteristics can provide useful information regarding the nature of the noise sources themselves as well as physical features within the oceans. Examples of inverse applications of the noise field include object imaging (‘acoustic daylight’), wind speed determination and rainfall measurements (see Chapter 6). The reverberation field in the oceans is the product of acoustic scattering by the surface and bottom boundaries, and by inhomogeneities within the oceans. The utility of the reverberation field as an inverse sensing technique is analogous to that of the ambient noise field. For example, the reverberation field can be inverted to image the sea floor (see Chapter 8). Table 1.1 summarizes selected inverse ocean acoustic-sensing techniques according to the natural phenomenon utilized. The specific techniques identified in Table 1.1 will be discussed in appropriate sections throughout this book. Table 1.1 Summary of inverrse ocean acoustic-sensirng techniques Propagation

Noise

Reverberation

• Tomography –density field (eddies, currents) –bulk temperature (climate monitoring)

• Surface noise –rainfall rates –wind speeds

• Field inversion –sea-floor imaging

MEASUREMENTS AND PREDICTION

Propagation

Noise

7

Reverberation

• Acoustic daylight –object imaging • Matched field processing –source localization –marine environment reconstruction • Bottom interaction –sea-floor composition –sea-floor scattering characteristics

Inverse acoustic-sensing techniques presently constitute adjuncts to direct measurement methods. However, the application of inverse acoustic-sensing techniques to dynamical studies of the oceans’ boundaries and interior show great promise for three reasons. First, such data can be used to establish comparative baselines for other remote sensors, such as satellites, by providing synoptic portraitures of the interior oceans together with concurrent groundtruth data at the sea surface. Second, inverse acousticsensing techniques often afford useful insights into a broad class of oceanic phenomena since their successful employment relies heavily on the use of numerical models to understand the role of the oceans as an acoustic medium. Third, inverse data provide estimates of spatially integrated and temporally averaged oceanic conditions that are not readily available from traditional oceanographic sensors. In furtherance of technology transfers from defense to commercial applications, proposals have been advanced to convert existing undersea surveillance assets into a National Acoustic Observatory. Possible research uses include (e.g. Amato, 1993; Carlson, 1994): • • • • •

stock and migration monitoring of large marine mammals remote ocean observations seismic and volcanic monitoring acoustic telemetry fisheries enforcement.

2 Acoustical oceanography

2.1 BACKGROUND Acoustical oceanography describes the role of the ocean as an acoustic medium. It relates oceanic properties to the behavior of underwater acoustic propagation, noise and reverberation. Acoustical oceanography crosses four other branches of oceanography: physical, chemical, geological and biological oceanography. The single most important acoustical variable in the ocean is sound speed. The distribution of sound speed in the ocean influences all other acoustic phenomena. The sound speed field in turn is determined by the density (or temperature and salinity) distribution in the ocean. Advection of the underwater sound field by water currents is also important. Refraction of sound by fronts, eddies and other dynamic features can distort signals. Knowledge of the state of the sea surface as well as the composition and topography of the sea floor is important for specification of boundary conditions. Biological organisms contribute to the noise field and also scatter underwater sound signals. The balance of this chapter addresses (1) physical and chemical properties, (2) sound speed, (3) boundaries, (4) dynamic features and (5) biologics. A number of books and published papers already exist on these subjects and appropriate citations will be made to them. Notable text and reference books of general nature include those by Apel (1987), Gill (1982) Clay and Medwin (1977), Neumann and Pierson (1966), Peixoto and Oort (1992), Pickard and Emery (1990) and Sverdrup et al. (1942). 2.2 PHYSICAL AND CHEMICAL PROPERTIES Temperature is basic to any physical description of the oceans. It is the easiest and therefore the most common type of oceanographic measurement made. The exchange of heat between the ocean and the atmosphere depends strongly on temperature. The density and resulting stratification of the ocean depend largely on temperature. The speed of sound in the upper layers of the ocean is most strongly dependent on temperature. Temperature further influences the kinds and rates of chemical reactions occurring in the ocean. The distribution of nutrients and other biologically important substances depends on temperature and the resulting density stratification. Sea water is a binary fluid in that it consists of various salts in water. Salinity is a term used to measure the quantity of salts dissolved in sea water and is expressed in units of parts per thousand (‰ or ppt). The precise definition of salinity is complicated. Fofonoff (1985) has reviewed the development of the modern

ACOUSTICAL OCEANOGRAPHY

9

salinity scale and the equation of state for sea water. The Practical Salinity Scale 1978 was introduced to rectify shortcomings associated with the traditional chlorinity-conductivity relationship used to establish salinity (Lewis, 1980; Perkin and Lewis, 1980; Culkin and Ridout, 1989). In the new scale, the existing link between chlorinity and salinity was broken in favor of a definitive salinity-conductivity relationship. The new practical standard is IAPSO (International Association for the Physical Sciences of the Ocean) Standard Seawater, produced and calibrated by the IAPSO Standard Seawater Service. Salinity is now a dimensionless quantity (psu, or practical salinity unit) because the algorithms in the new scale were adjusted to eliminate the ‰ (ppt) used in previous scales. The presence of salts affects a number of oceanic parameters including compressibility, sound speed, refractive index, thermal expansion, freezing point and temperature of maximum density. The density of sea water is related to temperature, salinity and pressure (which is nearly linearly proportional to depth) through the equation of state (e.g. Fofonoff, 1985). Density provides a measure of the hydrostatic stability in the ocean. Specifically, a stable water column is one in which density increases monotonically with depth. Sea water is compressible, although less so than pure water. The compressibility of sea water can be expressed by the coefficient of compressibility, which relates fractional changes in water volume to the corresponding changes in pressure (e.g. Apel, 1987). Compressibility of sea water is an important factor in several applications: the precise determination of the density of sea water, particularly at great depths; the computation of adiabatic temperature changes in the ocean (in an adiabatic process, compression results in warming, while expansion results in cooling); and most importantly for us, the computation of sound speed in sea water. The speed of sound (c) in sea water is related to the isothermal compressibility (K) as (2.1) where is the ratio of specific heats at constant pressure and constant volume, and is the density of sea water. The isothermal compressibility is easier to measure experimentally than is the adiabatic compressibility. 2.2.1 Temperature distribution The distribution of temperature at the surface of the oceans is zonal in nature, with isotherms (lines of constant temperature) oriented in an east-west pattern. The annual mean temperature distribution shown in Fig. 2.1 illustrates this general zonal gradation. This pattern is due largely to the zonal distribution of the solar energy received at the sea surface. Specific exceptions to this pattern occur in regions of upwelling (where colder water from below is brought to the surface through action of the winds), and in the vicinity of major current systems such as the Gulf Stream (where the temperature field is distorted). The relatively low equatorial and tropical sea-surface temperatures in the eastern Pacific and Atlantic Oceans are generally ascribed to upwelling. The more meridional trend of the isotherms off the northeast coast of the United States, for example, is evidence of the Gulf Stream current system. Examining only annual averages can sometimes be misleading. The monsoon circulation in the Indian Ocean, for example, makes interpretation of an annual mean temperature field questionable. The temperature field in the ocean exhibits a high degree of stratification with depth. Since the isotherms are nearly parallel to the horizontal plane, this type of structure is referred to as horizontal stratification.

10

PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.1 Annual mean temperature (°C) at the sea surface (Levitus, 1982). The distribution of surface temperatures shows a strong latitudinal dependence, due largely to the zonal distribution of solar energy received at the sea surface.

This is evidenced in Fig. 2.2, which presents zonal averages of temperature in the Atlantic Ocean by onedegree latitude belts. These zonal averages do not include the Mediterranean Sea, the Baltic Sea or Hudson Bay. 2.2.2 Salinity distribution The distribution of salinity at the surface of the ocean is shown in Fig. 2.3, and notable features have been summarized by Levitus (1982). Specifically, subtropical salinity maxima associated with the excess of evaporation over precipitation appear in all the individual oceans. Subpolar regions exhibit low salinities associated with the excess of precipitation over evaporation. Low-salinity tongues associated with runoff from the major river systems, such as the Amazon, are also apparent. Unlike the temperature fields, salinity does not exhibit a consistent stratification with depth in the Atlantic Ocean (Fig. 2.4). These patterns reflect the complex movements of water throughout the oceans. 2.2.3 Water masses Discriptions of sea-water characteristics and motions are facilitated by using the concept of water masses. This concept is analogous to that employed by meteorologists to describe air masses in weather patterns. Air masses are identified by characteristic combinations of air temperature and moisture content. These characteristics allow meteorologists to identify the past history (or source regions) of the various air masses. Examples of air masses include continental polar (cold, dry air formed over highlatitude land areas) and maritime tropical (warm, moist air formed over equatorial ocean areas). Oceanographers (e.g. Sverdrup et al., 1942) have convincingly demonstrated that certain characteristic combinations of water temperature and salinity are associated with water masses formed in particular regions of the world’s oceans. These water masses then spread both vertically and laterally and occupy depth ranges of the water column consistent with their density and are distinguishable from one another

ACOUSTICAL OCEANOGRAPHY

11

Fig. 2.2 Annual mean Atlantic Ocean zonal average (by one-degree squares) of temperature (°C) as a function of depth (Levitus, 1982). (Note break in the depth scale at 1000m.)

when plotted on a graph of temperature versus salinity, referred to as a T-S diagram. The T-S relations of the principal water masses of the Atlantic Ocean are presented in Fig. 2.5 (Naval Oceanographic Office, 1972). Approximate water depths corresponding to the occurrence of particular T–S combinations are also indicated. Emery and Meincke (1986) have provided an updated review and summary of the global water masses. The movement of Antarctic Intermediate Water (AAIW), for example, can now be identified in Fig. 2.4. Specifically, the movement of AAIW is evidenced by a low-salinity tongue (34.0–34.6 ppt) extending downward in the latitude belt 60°S to 70°S, then northward at a depth of about 700–800 m, and finally upward near the equator. The distribution of sound speed in the ocean can be related to the local water mass structure. Knowledge of the water mass structure, then, can greatly enhance our understanding of the large-scale spatial and temporal variability of the sound speed field in the ocean. 2.3 SOUND SPEED The speed of sound in sea water is an oceanographic variable that determines the behavior of sound propagation in the ocean. Sound speed varies as a function of water temperature, salinity and pressure (or depth). The speed of sound in sea water increases with an increase in any of these three parameters. The term ‘sound velocity’ is an older term that is used synonymously with ‘sound speed’.

12

PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.3 Annual mean salinity (ppt) at the sea surface (Levitus, 1982).

2.3.1 Calculation and measurements Many empirical relationships have been developed over the years for calculating sound speed using values of water temperature, salinity and pressure (or depth). Frequently used formulas include those of Wilson (1960), Leroy (1969), Frye and Pugh (1971), Del Grosso (1974), Medwin (1975), Chen and Millero (1977), Lovett (1978), Coppens (1981) and Mackenzie (1981). Each formula has its own limits for temperature, salinity and pressure (or depth). Calculations outside of the specified limits may be in error. Recent tomographic measurements of sound speed in the oceans have implications for very-long-range ducted propagation (Spiesberger and Metzger, 1991b). Specifically, previous algorithms derived from laboratory measurements have been found to overpredict sound speeds due to pressure effects at great depths (Dushaw et al., 1993). This matter is still the subject of investigation. For convenience, the formula developed by Mackenzie (1981) is presented here: (2.2)

where c is the speed of sound in sea water (in meters per second), T is the water temperature (in degrees Celsius), S is the salinity (in parts per thousand) and D is the depth (in meters). This equation is valid over the following ranges: 0°C T 30°C; 30 ppt S 40 ppt; 0m D 8000m. Mackenzie also discussed the relationship between pressure and depth in the ocean. Two practical devices commonly used to measure sound speed as a function of depth in the ocean are the bathythermograph (BT) and the velocimeter. The bathythermograph, which is oftan an expendable device (designated XBT), actually measures temperature as a function of depth. Sound speed can then be calculated using one of the available empirical formulas, often on the assumption that salinity is a constant, or nearly so. This assumption is justified by the observation that the typical range of salinities in the open ocean is usually small and that the corresponding impact on sound speed is negligible from a practical standpoint. In coastal areas, and near rivers or ice, this assumption is not generally valid.

ACOUSTICAL OCEANOGRAPHY

13

Fig. 2.4 Annual mean Atlantic Ocean zonal average (by one-degree squares) of salinity (ppt) as a function of depth (Levitus, 1982). (Note break in the depth scale at 1000m.)

The second device, a velocimeter, measures sound speed directly in terms of the travel time of sound over a constant-length fixed path. An expendable version of this instrument is also available (designated XSV). The velocimeter would seem to be preferred instrument for obtaining measurements of sound speed in support of naval operations. From a broader scientific perspective, however, information on the temperature distribution in the ocean directly supports many other naval applications including ocean dynamics modeling, air–sea interaction studies and various marine biological investigations, to name a few. The demonstrated accuracy of the various empirical formulas (when used within the specified limits) is sufficient for most operational applications. In terms of expendable sensors, measurements of water temperature using the bathythermograph represent a better scientific investment. Nevertheless, some naval operations in coastal areas or in ice-covered regions may be better supported by the velocimeter. Increasing emphasis on naval operations in the littoral zone has placed greater importance on measuring, in real time, coastal processes that are characterized by extreme temporal and spatial variations. Precise scientific measurements are normally accomplished using ex-pensive, but recoverable, instrumentation. Such instruments are typically deployed from surface ships which maintain station while the instrument package is lowered and then raised in what is termed a hydrographic cast. The instrument package can contain an assortment of sensors for measuring water temperature, salinity (or conductivity), pressure, sound speed, currents and dissolved oxygen, among others.

14

PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.5 Temperature-salinity (T–S) diagrams of the major water masses of the Atlantic Ocean (Naval Oceanographic Office, 1972).

2.3.2 Sound speed distribution Typical North Atlantic winter and summer profiles of sound speed versus depth are shown in Fig. 2.6. These profiles represent a region of the North Atlantic Ocean located near 23°N and 70°W (Naval Oceanographic Office, 1972). Temperature–salinity (T–S) diagrams for both winter and summer seasons, based on actual measurements, are also presented to show their relationships to the sound speed profiles. Since the T–S diagrams indicate the ocean depths corresponding to the measured temperature–salinity pairs, the individual temperature and salinity profiles for both winter and summer can be reconstructed. Principal water masses are noted on the sound speed profiles. The so-called 18° Water marks a change in the sound speed gradient at a depth of about 300 m, and Mediterranean intermediate water (MIW) occupies the region of the water column near the sound speed minimum (at about 1200m). The profiles in Fig. 2.6 are representative of the sound speed profiles encountered in many tropical and subtropical ocean areas. Such profiles may be divided into arbitrary layers, each having different characteristics and occurrence (Fig. 2.7). Just below the sea surface is the sonic layer where the speed of sound is influenced by local changes in heating, cooling and wind action. The base of the sonic layer is defined as the sonic layer depth (SLD), which is associated with the near-surface maximum in sound speed. This surface layer is usually associated with a well-mixed layer of isothermal water. Oceanographers refer to this well-mixed region as the mixed layer; the base of this layer is then termed the mixed layer depth (MLD).

ACOUSTICAL OCEANOGRAPHY

15

Below the mixed layer lies the thermocline, a region of the water column in which the temperature decreases rapidly with depth. This region of the water column is also characterized by a negative sound speed gradient (i.e. sound speed decreases with depth). Below the thermocline and extending to the sea floor is the deep isothermal layer. This layer has a nearly constant temperature in which the speed of sound increases with depth because of the effect of pressure on sound speed. In this region, the sound speed profile becomes linear with a positive gradient of about 0.017 (m/s)/m or 0.017/s. Between the negative sound speed gradient of the thermocline and the positive gradient of the deep isothermal layer is a sound speed minimum. The depth corresponding to this sound speed minimum is referred to as the sound channel axis. At high latitudes, the deep isothermal layer extends nearly to the sea surface; that is, the sound channel axis shoals as one approaches the polar regions. This behavior is vividly demonstrated in Fig. 2.8; the top panel presents contours of the depth of the minimum sound speed (or channel) axis (m) and the bottom panel presents the sound speed (m/s) on this axial surface (Munk and Forbes, 1989). At low latitudes, the depth of the sound channel axis is typically near 1000 m; at high latitudes, the axis is located near the sea surface. The associated sound speeds on this axial surface generally decrease away from the equatorial regions. In profiles containing a sound channel axis, a critical depth can be defined as that depth below the axis at which the sound speed equals the near-surface maximum value. (The near-surface maximum value of sound speed is usually located at the sonic layer depth). The vertical distance between the critical depth and the sea floor is referred to as the depth excess. Other pairs of points can be identified on the sound speed profile which have the same value of sound speed but lie on opposite sides of the sound channel axis; such pairs are referred to as conjugate depths. In Fig. 2.7, the critical depth is actually a conjugate of the sonic layer depth. Additional sound speed profiles typical of the winter season in different ocean areas of the world are presented in Fig. 2.9 to demonstrate that the simple model for sound speed profiles described above is not applicable to high-latitude ocean areas or to some smaller water bodies. The depth dependence of sound speed in the ocean poses a particular problem for echo sounders which use near-vertical acoustic paths to measure the depth of the sea floor based on the two-way travel time of the signal. Echo sounders are set to read the depth directly by assuming a constant speed of sound in the water column, usually 1463 m/s or 1500 m/s. When the depth-integrated (or mean) sound speed departs from the assumed value, a correction must be applied to the observed readings. Bialek (1966, p. 63), for example, has tabulated such corrections according to ocean area. Depending on the particular ocean area and water depth, these corrections can be on the order of several percent of the true water depth. Underwater acoustic propagation problems involving long range may not be able to ignore the hotizontal variations in either sound speed or bathymetry. Modeling developments, therefore, generally distinguish between range-independent (the ocean varies only as a function of depth) and range-dependent (the ocean varies as a function of both depth and range) problems. Parameters other than sound speed and water depth may also be considered in range-dependent problems: surface losses, bottom losses and absorption, for example. An east–west cross-section of the North Atlantic Ocean between 23°N and 24°N is presented in Fig. 2.10 to illustrate the variability of both sound speed and bathymetry over moderate range scales. Also noteworthy are changes in the depth of the sound channel axis and the absence of a critical (or limiting) depth in some basins. The sound speed profiles presented previously in Fig. 2.6 are consistent with the data of Fig. 2.10 at a longitude of about 70°W.

Fig. 2.6 Sound speed profiles (winter and summer) and T–S comparisons for the North Atlantic Ocean near 23ºN, 70ºW (Naval Oceanographic office, 1972).

16 PHYSICAL AND CHEMICAL PROPERTIES

ACOUSTICAL OCEANOGRAPHY

17

Fig. 2.7 Schematic relationship between temperature and sound speed profiles in the deep ocean.

2.4 BOUNDARIES The boundaries of the water column—the sea surface and the sea floor— can exert a profound influence on the propagation of acoustic energy through the action of reflection, scattering and absorption. 2.4.1 Sea surface The surface of the sea is both a reflector and a scatterer of sound. If the sea surface were perfectly smooth, it would form an almost perfect reflector of sound due to the acoustic impedance mismatch at the air–water interface. As the sea surface becomes rough, as it does under the influence of wind (see Table 2.1), reflection losses are no longer near zero. Sea-surface roughness is typically specified in terms of wave height. Observations of weather at sea generally record wind speed and not wave height as a description of sea state (Bowditch, 1977). A number of statistical relationships exist with which to quantitatively associate these two parameters (e.g. Earle and Bishop, 1984). These relationships are very precise as to the height above sea level at which wind speed is observed and the type of statistical wave height considered. Based on the Pierson–Moskowitz spectrum (Moskowitz, 1964; Pierson and Moskowitz, 1964; Pierson, 1964, 1991), a fully developed significant wave height can be calculated from the observed wind speed as (2.3)

18

PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.8 Global maps of the sound channel axis. Upper panel: channel depth (m). Bottom panel: channel sound speed (m/ s) (Munk and Forbes, 1989; J. Phys. Oceanogr., 19, 1765–78; copyright by the American Meteorological Society).

ACOUSTICAL OCEANOGRAPHY 19

Fig. 2.9 Characteristic winter sound speed profiles for selected deep-ocean areas of the world.

20 PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.10 East—west sound speed cross-section between 23ºN and 24ºN and between 60ºW and 100ºW for the period February— April (Naval Oceanographic office, 1972).

ACOUSTICAL OCEANOGRAPHY 21

22

PHYSICAL AND CHEMICAL PROPERTIES

Table 2.1 Beaufort Wind Scale with corresponding Sea State Codes (Bowditch, 1977) Beaufort Wind speed number or force

World Estimating wind speed Meteorological Organizat ion (1964)

Sea state

Knots

mph

Meters per second

km per hour

Effects observed far from land

Effects Effects observed observed near coast on land

Term and Code height of waves, in meters

0

Under 1

Under 1

0.0–0.2

Under 1

Calm

Sea like mirror

Calm

1

1–3

1–3

0.3–1.5

1–5

Light air

Fishing smack just has steerage way

2

4–6

4–7

1.6–3.3

6–11

Light breeze

Ripples with appearan ce of scales; no foam crests Small wavelets; crests of glassy appearan ce, not breaking

3

7–10

8–12

3.4–5.4

12–19

Gentle breeze

4

11–16

13–18

5.5–7.9

20–28

Moderat e breeze

5

17–21

19–24

8.0–10. 7

29–38

Fresh breeze

Large wavelets; crests begin to break; scattered whitecap s Small waves, becomin g longer; numerou s whitecap s Moderat e waves, taking longer

Wind fills the sails of smacks which then travel at about 1– 2 miles per hour Smacks begin to careen and travel about 3– 4 miles per hour Good working breeze, smacks carry all canvas with good list Smacks shorten sail

Calm; smoke rises vertically Smoke drift indicates wind direction; vanes do not move Wind felt on face; leaves rustle; vanes begin to move

Leaves, small twigs in constant motion; light flags extended Dust, leaves and loose paper raised up; small branches move Small trees in leaf

Calm, glassy, 0

0

Calm, rippled 0–0.1

1

Smooth, wavelets, 0.1–0.5

2

Slight, 0. 5–1.25

3

Moderat e, 1.25– 2.5

4

ACOUSTICAL OCEANOGRAPHY

Beaufort Wind speed number or force

Knots

mph

Meters per second

World Estimating wind speed Meteorological Organizat ion (1964) km per hour

Effects observed far from land

Effects Effects observed observed near coast on land form; many whitecap s; some spray Larger waves forming; whitecap s everywh ere; more spray

6

22–27

25–31

10.8– 13.8

39–49

Strong breeze

7

28–33

32–38

13.9– 17.1

50–61

Near gale Sea heaps up; white foam from breaking waves begins to be blown in streaks

8

34–40

39–46

17.2–20. 7

62–74

Gale

9

41–47

47–54

20.8–24. 4

75–88

Strong gale

Moderatel y high waves of greater length; edges of crests begin to break into spindrift; foam is blown in wellmarked streaks High waves;

23

Sea state

Term and Code height of waves, in meters begin to sway

Smacks have doubled reef in mainsail; care required when fishing Smacks remain in harbor and those at sea lieto

All smacks make for harbor, if near

Larger branches of trees in motion; whistling heard in wires

Rough 2. 5–4

5

Whole trees in motion; resistanc e felt in walking against wind

Twigs and small branches broken off trees; progress generally impeded

Slight structural

Very rough, 4– 6

6

24

PHYSICAL AND CHEMICAL PROPERTIES

10

48–55

55–63

24.5–28. 4

89–102

Storm

11

56–63

64–72

28.5–32. 6

103–117

Violent storm

12 64 and over

73 and over

32.7 and over

118 and over

Hurricane

sea begins to roll; dense streaks of foam; spray may reduce visibility Very high waves with overhanging crests; sea takes white appearanc e as foam is blown in very dense streaks; rolling is heavy and visibility reduced Exception ally high waves; sea covered with white foam patches; visibility still more reduced Air filled with foam; sea completel y white with driving spray; visibility greately reduced

damage occurs; slate blown from roofs

Seldon experienc ed on land; trees broken or uprooted; considera ble structural damage occurs

High, 6–9

7

Very rarely experienc ed on land; usually accompa nied by widespre ad damage

Very high, 9– 14

8

Phenome nal, over 14

9

where V is the wind speed (in knots), measured at a height of 19.5 m, and H1/3 is the average height (in meters) of the one-third highest waves.

ACOUSTICAL OCEANOGRAPHY

25

Two other commonly used wave height descriptors are the rms wave height (Hrms) and the one-tenth significant wave height (H1/10). These are related to the one-third significant wave height as (2.4) (2.5) These relationships can be compared with those presented in Table 2.1, which are based on the rms wave heights. It should be evident that care is needed when specifying the type of wave height being considered in the calculations. Air bubbles are produced by the breaking of waves and are carried by turbulence beneath the surface. They are also generated in the wakes of ships where they persist for long periods of time. Free air bubbles in the sea are quite small since the larger bubbles tend to rise quickly to the surface. Bubbles only form a very small volumetric percentage of the sea; however, because air has a markedly different density and compressibility from that of sea water and because of the resonant characteristics of bubbles, the suspended air content of sea water has a profound effect upon underwater sound. Urick (1983, pp. 249–54) has summarized these effects which include resonance and changes in sound speed. Aside from reflection losses, there are other acoustic effects associated with interactions with the sea surface. A moving sea surface produces frequency-smearing and shifting effects on constant-frequency signals. Large and rapid fluctuations in amplitude or intensity are also produced by reflection at the sea surface. Furthermore, Lloyd mirror, or image-interference, effects produce a pattern of constructive and destructive interference between direct and surface-reflected sound. When the sea surface is roughened by wind, this effect is lessened. 2.4.2 Ice cover When the sea surface is covered by an ice canopy, as in the polar regions (see Figs 2.11 and 2.12), acoustic interaction with the surface is further complicated by an irregular under-ice surface. The Arctic environment can be segregated into three distinct regions according to the type of ice cover: (1) pack ice, (2) Marginal Ice Zone (MIZ) and (3) open ocean. The differing geometrical configurations associated with the degree of ice cover and depth of water afford substantially different environmental impacts on acoustic system operation and performance. Field measurements have shown that forward scatter from a rough anisotropic ice canopy is a function of acoustic frequency, geometry and the statistical (spatial correlation) properties of the under-ice surface. 2.4.3 Sea floor The sea floor is a reflecting and scattering boundary having a number of characteristics similar in nature to those of the sea surface. Its effects, however, are more complicated than those of the sea surface because of its diverse and multilayered composition. Specifically, the sea floor is often layered, with a density and sound speed that may change gradually or abruptly with depth or even over short ranges. Furthermore, the sea floor is more variable in its acoustic properties since its composition may vary from hard rock to soft mud. One feature that is distinct from the sea surface is that the bottom characteristics can be considered to be constant over time, whereas the configuration of the sea surface is statistically in a state of change as the wind velocity changes.

26

PHYSICAL AND CHEMICAL PROPERTIES

Fig. 2.11 Average boundaries of sea ice (coverage at least 5–8 tenths) in autumn and spring in the Arctic. Arrows indicate the general drift pattern. The width of the stippled area indicates the range of ice limits between autumn and spring (Untersteiner, 1966).

Because of the variable stratification of the bottom sediments in many areas, sound is often transmitted into the bottom and refracted or reflected internally. Thus, the bottom often becomes a propagating medium characterized by both shear and compressional sound speeds. The topography of the sea floor exhibits a diversity of features not unlike those of the continental land masses. Figure 2.13 presents an artist’s conception of common ocean basin features. The undersea features noted in Fig. 2.13 are defined in the Glossary of terms (Appendix B). Underwater ridges and seamounts can effectively block the propagation of sound. When actively ensonified, seamounts can mask targets of interest by either providing false targets or by shadowing targets of interest. Marine seismic studies have greatly improved our understanding of the crustal structure of that part of the earth covered by the oceans (Bryan, 1967); a typical bottom structure section is presented in Fig. 2.14. This typical section consists of 5–6 km of water, about 0.5 km of unconsolidated sediments, 1–2 km of basement rock and 4–6 km of crustal rock overlying the upper mantle. Major portions of the sea floor are covered with unconsolidated sediments with an average thickness of approximately 500 m. Sediments can be classified according to their origin as either terrigenous or pelagic, although no single classification scheme has universal approval. The general distribution of sediments according to Arrhenius (1963) is presented in Fig. 2.15. Terrigenous sediments are derived from land and are particularly prominent near the mouths of large rivers. These sediments are generally classified as silt, sand and mud.

ACOUSTICAL OCEANOGRAPHY

27

Fig. 2.12 Average boundaries of sea ice (coverage at least 5–8 tenths) in autumn and spring in the Antarctic. Arrows indicate the general drift pattern. The width of the stippled area indicates the range of ice limits between autumn and spring (Untersteiner, 1966).

Pelagic sediments are derived from two sources: organic and inorganic. Organic sediments comprise the remains of dead organisms and are classified as either calcareous or siliceous oozes. Inorganic pelagic sediments are derived from materials suspended in the atmosphere and are generally classified as clay. 2.5 DYNAMIC FEATURES It is convenient to categorize dynamic features of the ocean according to characteristic time and space scales. While no precise terminology is universally accepted, it is common to recognize space scales as large (>100 km), meso (100m–100 km) and fine (30 m) are less likely to be affected by internal waves. Above 20 kHz (with acoustic wavelengths less than a few centimeters), the effects of fine-scale features are probably more important (Spindel, 1985). Zhou et al. (1991) investigated the interaction of underwater sound with internal gravity waves in an attempt to explain the anomalous behavior of low-frequency (~ 300–1100 Hz) acoustic propagation conditions observed in some shallow-water areas. As a result of this investigation, it was noted that acoustic measurements could be employed in an inverse fashion for the remote sensing of internal wave activity in the coastal zone. Chin–Bing et al. (1993a) numerically simulated low-frequency backscatter from internal waves in shallow water. One problem in testing theories of acoustic propagation in inhomogeneous media is the inability to determine accurately the spatial and temporal scales associated with fluctuations in the index of refraction. These fluctuations can be caused by tides, internal waves and fine-scale features. Ewart and Reynolds (1984) reported results from the Mid-Ocean Acoustic Transmission Experiment (MATE). This experiment was designed to measure phase and intensity fluctuations in sound pulses transmitted at 2, 4, 8 and 13 kHz over a wholly refracted path (i.e. no boundary interactions). Two receiver towers were placed on Cobb Seamount (in the northeastern portion of the North Pacific Ocean); a sister tower located 20 km to the southwest was used for placement of the acoustic transmitter tower. This geometry minimized transducer motion and also assured the presence of a wholly refracted path between the source and receiver. The environmental program conducted in support of MATE was specifically designed to oversample the internal-wave variability within the context of the Garrett–Munk model (Garrett and Munk, 1979). These environmental measurements can also be used to test future models of internal-wave variability. 2.5.3 Fine-scale features One type of fine-scale oceanic feature is the ‘thermohaline staircase’. These staircases are generally found in the main thermocline and are evidenced by layers of uniform temperature and salinity on the order of 10

ACOUSTICAL OCEANOGRAPHY

39

Fig. 2.24 Series of expendable bathythermograph (XBT) temperature profiles from an east–west section east of Barbados in the Atlantic Ocean. The profiles are separated by a distance of 5.5 km; the total distance covered is 220 km. The temperature scale is correct for the profile at the extreme left (west), and each subsequent profile is offset by 1. 6°C Undulations of the thermocline caused by internal waves and meso-scale eddies can also be seen (Schmitt, 1987; EOS, Trans. Amer. Geophys. Union, 68, 57–60; copyright by the American Geophysical Union).

m thick separated by thin, high-gradient interfaces on the order of a few meters in thickness (Fig. 2.24). The incidence of well-developed staircases appears to be limited to less than 10% of the available highresolution profiles taken in the North Atlantic Ocean (Schmitt, 1987; Schmitt et al., 1987). Staircase structures are most frequently associated with a strongly destabilizing vertical gradient of salinity. These features have been observed northeast of South America in the tropical Atlantic Ocean (just outside the Caribbean Sea), and also in the eastern Atlantic outside the Mediterranean Sea. Before the dynamics of the thermohaline staircases were understood, these features were sometimes regarded as malfunctions in the oceanographic sensors which recorded them. Research is not conclusive as to what effect these staircase features might have on underwater acoustic propagation. Chin–Bing et al. (1994) studied the effects of thermohaline staircases on low-frequency (50 Hz) sound propagation. Several propagation models were used to generate transmission loss as a function of range from source to receiver using a sound speed profile containing staircase features. A source was placed (in depth) at the center of the staircase features while receivers were placed above, below, and at the center of the features. These results were then compared to baseline simulations using a profile in which the effects of the staircase features were effectively averaged out. The greatest effects were observed when both the source and receiver were placed at the center of the features. These effects were attributed to a redistribution of intensity caused by the staircase features. Chin–Bing et al. (1994) also noted that backscatter can occur when the step-structured discontinuities of the thermohaline staircases are on the order of an acoustic wavelength; thus, at frequencies greater than about 3 kHz, backscatter from the thermohaline steps could be significant. 2.6 BIOLOGICS Marine organisms can simplistically be segregated into four major categories: plankton, nekton, benthos and algae. Plankton (or floaters) include both plants (phytoplankton) and animals (zooplankton). The zooplankton have little or no swimming ability and thus drift with the currents. Phytoplankton are typically smaller than 0.5 mm while zooplankton are smaller than 1 cm.

40

PHYSICAL AND CHEMICAL PROPERTIES

Nekton (or free swimmers) are animals which are capable of swimming purposefully. Nekton include fish and mammals, and occur over the entire depth range of the ocean. Benthos are dwellers on, in or near the bottom. Fouling organisms such as barnacles would also be included in this category. Algae include marine plant life, such as seaweed. Biological organisms can affect underwater sound in several ways: through noise production; attenuation and scattering of signals; presentation of false targets; and fouling of sonar transducers. Certain marine animals, many of which are found over the continental shelves, produce sounds that increase the background (or ambient) noise levels: snapping shrimp, whales, porpoises, and various fish such as croakers and drum fish. Organisms that may cause attenuation are schools of fish, dense populations of plankton and floating kelp, for example. False targets are commonly presented to active sonars by whales or large schools of fish or porpoises. While fouling organisms such as barnacles do not directly affect sound, indirectly they can degrade sonar performance by fouling sonar domes and transducer faces. Furthermore, such organisms can contribute to an increase in hull noise of ships and submarines through the generation of turbulence as the vessels move through the water. This effect is also referred to as self noise. Perhaps the most notable impact of marine organisms on active sonars (particularly those operating at frequencies near 10 kHz) is known as the deep scattering layer (DSL). The DSL is a dense accumulation of marine organisms at depth below the surface. The strong scattering nature of the DSL is primarily attributable to fish and other marine animals with swim bladders and gas floats, although plankton and nekton are also present. The DSL is typically encountered in temperate regions and migrates vertically, being near the surface at night and at depths near 1000 m during the day.

3 Propagation I. Observations and physical models

3.1 BACKGROUND The propagation of sound in the sea has been intensely studied since the beginning of World War II when it was recognized that an understanding of this phenomenon was essential to the successful conduct of AntiSubmarine Warfare (ASW) operations. These early measurements were quickly transformed into effective, albeit primitive, prediction tools. ASW has continued to be the primary motivation behind modern advances in propagation modeling. The study of sound propagation in the sea is fundamental to the understanding of all other underwater acoustic phenomena. The essentiality of propagation models is thus inherent in the hierarchy of acoustic models, as illustrated previously in Fig. 1.1. Advances in propagation modeling have been achieved by both marine seismologists and underwater acousticians, although the motivating factors have been quite different. Marine seismologists have traditionally used earthborne propagation of elastic waves to study the solid earth beneath the oceans. Underwater acousticians have concentrated on the study of waterborne, compressional-wave propagation phenomena in the ocean as well as in the shallow subbottom layers (Akal and Berkson, 1986). In recent years, research in underwater acoustics has been extended to frequencies below several hundred hertz. This overlaps with the spectral domain of marine seismologists. Moreover, marine seismologists have become more interested in exploring the velocity–depth structure of the uppermost layers of the sea floor using higher frequencies. This area of overlapping interests has been recognized as a subdiscipline of both communities and is referred to as ‘ocean seismo-acoustics’. The emphasis in this chapter is focused on underwater acoustic applications. Developments in marine seismology will be mentioned when the applications to sonar modeling are clearly evident. Much research has been performed in the marine seismology community that is theoretically and conceptually applicable to underwater acoustics. Such practical research includes the development of sophisticated yet robust mathematical methods. Propagation models have continued to be used for the prediction of sonar performance. They have also found great utility in analyzing field measurements, in designing improved sonar systems, and in designing efficient field experiments. As modeling has become prominent in many aspects of underwater acoustics, it was deemed timely to take stock of the state of the art in modeling techniques and their relationship to available measurements. Ideally, such an assessment should identify those areas requiring further measurement support as well as those that are firmly understood and hence properly-modeled.

42

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.1 Example of a simple experimental geometry (adapted from Ingenito et al., 1978).

This chapter addresses the observations that have been made in the field and the physical models that have been developed. Aspects of propagation phenomena—ducts and channels, boundary interactions, volumetric effects and coherence—are described. Chapter 4 will address the mathematical models that have been developed for underwater acoustic propagation. Specialized aspects of surface ducts, shallow water areas and Arctic regions are discussed in Chapter 5. 3.2 NATURE OF MEASUREMENTS Field measurement programs are usually quite complex and typically involve multiple platforms (e.g. ships, buoys, towers, aircraft, submarines or satellites). A wide variety of experimental field techniques have been used in underwater acoustic propagation studies. Some of the more typical types of measurement platforms and experimental geometries that have been utilized include (Urick, 1982, Chapter 1): 1. Two ships—one a source ship and the other a receiving ship; the range between them is changed as transmission runs are made in order to yield level versus range. 2. Single ship—using a suspended transmitter and either sonobuoys or a hydrophone array for reception. 3. Ship and aircraft—where the aircraft drops explosive sound sources while flying toward or away from the ship. 4. Single aircraft—using sonobuoys for reception and recording on board the aircraft. 5. Bottomed hydrophone array—with a cable connected to shore, receives signals transmitted from a ship or the explosive shots dropped by an aircraft. 6. Two bottomed transducers—one acting as a source and the other as a receiver; this geometry is typically used in studies of the fluctuation of sound transmission between two fixed points in the sea. A simple experimental geometry illustrating method (2) above is presented in Fig. 3.1. Here, the transmitting ship is receiving the signals via radio directly from the array.

BASIC CONCEPTS

43

Fig. 3.2 Example of standard transmission loss curves generated by the FACT model for each combination of frequency and source-receiver geometry. Here, the source and receiver depths are fixed at 150 m and 90 m, respectively. The peaks (minimum TL values) correspond to convergence zones. Note the increase in transmission loss with increasing frequency due to absorption.

Fully integrated oceanographic and acoustic field experiments are required in order to achieve a comprehensive portraiture of the temporal, spatial and spectral scales necessary for a full understanding of the governing acoustic phenomena. 3.3 BASIC CONCEPTS The standard unit of measure of underwater acoustic propagation is acoustic intensity (I), which is sound pressure flow (power) per unit area: (3.1) where p is the pressure amplitude, is the density of sea water and c the sound speed in sea water. The product c is commonly referred to as the acoustic impedance. Transmission loss (TL) can be defined as ten times the log (base 10) of the ratio of the reference intensity (Iref) measured at a point 1 m from the source, to the intensity (I), measured at a distant point, and is expressed in units of decibels (dB): (3.2) The standard metric unit for pressure (force per unit area) is 1 microPascal, which is equivalent to 10−6 newton/m2, and is abbreviated µPa. Transmission loss (TL) has conventionally been plotted for each frequency and source-receiver geometry as a function of range, as illustrated in Fig. 3.2. This type of display is easily generated by all propagation

44

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.3 Example showing contours of transmission loss plotted in the range– depth plane. This plot is valid for one frequency (30 Hz) and one source depth (50 m), but can be used to determine the transmission loss at any receiver location in the range–depth plane; the contour interval is 6dB (Schmidt, 1988).

models. Certain types of propagation models can also generate and display acoustic transmission loss in the entire range-depth plane for all receiver positions, given a fixed source depth (Fig. 3.3). Sonar performance is commonly described in terms of a figure of merit (FOM). The FOM is a quantitative measure of sonar performance: the larger the value, the greater the performance potential. Numerically, the FOM is equal to the allowable one-way transmission loss in passive sonars. The FOM is further described in Chapter 10 within the context of the sonar equations. The display method illustrated in Fig. 3.2 is very useful in evaluating passive sonar performance. Specifically, once the FOM is calculated for a particular sonar, a horizontal line can be drawn on the plot equating the numerical value of the FOM to transmission loss. Then, any area below the transmission loss curve, but above the FOM line, represents a sonar detection area. Figure 3.4 shows a hypothetical relationship between the FOM and the sonar detection areas, and the correspondence between the TL curve and the ray paths as propagated in the water column. Sound propagates in the sea by way of a variety of paths. The particular paths traveled depend upon the sound speed structure in the water column and the source–receiver geometry. These paths are: direct path, surface duct, bottom bounce, convergence zone, deep sound channel and reliable acoustic path. These six basic paths are illustrated in Fig. 3.5. Various combinations of paths are possible and are referred to as multipath propagation. In basic ray tracing, Snell’s law is used in one form or another. This law describes the refraction of sound rays in a medium in which sound speed varies as a function of depth, but is constant within discrete horizontal layers of the water column. Consider Fig. 3.6 where a ray (normal to the acoustic wavefronts) is traveling from medium 1 (with sound speed c1) into medium 2 (with sound speed c2), where c1 c2. Let 1 be the distance between successive wavefronts (wavelength) in medium 1 and 2 the corresponding value in medium 2. Then, as defined in Fig. 3.6 where t is an increment of time. Rearranging terms, we obtain the familiar relation (3.3a)

BASIC CONCEPTS

45

Fig. 3.4 Hypothetical relationship between (a) transmission loss (TL) curve and (b) the corresponding propagation paths and detection zones (cross-hatched areas near the surface) associated with a Figure of Merit (FOM) of 85 dB. A plausible sound speed profile is shown at the left of panel (b); both the source (target) and receiver (sonar) are positioned near the surface.

or equivalently, from Fig. 3.6, (3.3b) As a matter of convention, is referred to as the incidence angle while is referred to as the grazing angle. In a homogeneous medium, acoustic transmission loss varies as the inverse of the range squared. This relationship is easily derived as follows: Let I=intensity, P=power and A=area. Then

For spherical spreading (see Fig. 3.7(a)) where r1, r2 are radii of concentric spherical sections.

46

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.5 Six basic propagation paths in the sea: A, direct path (DP); B, surface duct (SD); C, bottom bounce (BB); D, convergence zone (CZ); E, deep sound channel (DSC); and F, reliable acoustic path (RAP).

Power (P) is conserved; therefore, P=I1A1=I2A2 and but r1=unit reference distance and thus Since intensity is power per unit area, and since the area of a sphere increases as the square of its radius, the intensity falls off as the inverse square of the radius (or range) in order that power remains constant. The corresponding transmission loss (TL) is defined as (3.4) This relationship would be valid for an isotropic deep ocean with no absorption effects. An analogous expression can be derived for cylindrical spreading. This spreading law would be appropriate for a duct, or in shallow water, where the water is homogeneous and the boundaries are perfect reflectors. Then (referring to Fig. 3.7(b)) where H is the depth of the duct or of the water column. Since power is conserved

For unit radius r1

BASIC CONCEPTS

47

Fig. 3.6 Geometry for Snell’s law.

Thus, the intensity falls off as the inverse of the radius (or range). The corresponding transmission loss (TL) is defined as (3.5) Equations (3.4) and (3.5) are valid in isotropic ocean environments. 3.4 SEA-SURFACE BOUNDARY The sea surface affects underwater sound by providing a mechanism for: 1. 2. 3. 4. 5.

forward scattering and reflection loss image interference and frequency effects attenuation by a layer of bubbles noise generation at higher frequencies due to surface weather backscattering and surface reverberation.

Urick (1982, Chapter 10) provides a comprehensive summary of sound reflection and scattering by the sea surface. Items (1) through (3) will be discussed below. Item (4) will be discussed in Chapter 6 and item (5) will be discussed in Chapter 8. The mechanisms operating at the surface can be incorporated into mathematical models through the specification of ‘boundary conditions’. These boundary conditions can range from simplistic to complex, depending upon the sophistication of the model and the availability of information concerning the state of the sea surface.

48

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.7 Geometry for (a) spherical spreading and (b) cylindrical spreading (Urick, 1983; Principles of Underwater Sound, 3rd edn; reproduced with permission of McGraw-Hill Publishing Company).

3.4.1 Forward scattering and reflection loss When a plane sound wave in water strikes a perfectly smooth surface, nearly all of the energy is reflected at the boundary in the forward (or specular) direction as a coherent plane wave. As the sea surface roughens under the influence of wind, sound is also scattered in the backward and out-of-plane directions, and the intensity in the forward direction is accordingly reduced. The backward-directed (backscattered) energy gives rise to surface reverberation. Eckart (1953) developed a theoretical treatment of scattering by a sinusoidal boundary as a way to approximate reflection from a wind-roughened sea surface. Marsh et al. (1961) developed simple formulas to express scattering losses at the sea surface. The sea surface is commonly modeled as a pressure release surface (see Kinsler et al., 1982, pp. 126–7). This is a condition in which the acoustic pressure at the air-water interface is nearly zero, the amplitude of the reflected wave (in water) is almost equal to that of the incident wave, and there is a 180° phase shift. It is also common practice to use the term ‘reflection coefficient’ to express the amount of acoustic energy reflected from a surface or boundary between two media. This coefficient depends upon the grazing angle and the difference in the acoustic impedance between the two media. A reflection loss is then defined as 10 log (reflection coefficient). This reflection loss is referred to as surface loss when describing the reflection of sound from the sea surface, or as bottom loss when describing the reflection of sound from the sea floor.

BASIC CONCEPTS

49

A measure of the acoustic roughness of the sea surface is provided by the Rayleigh parameter R through the relation (3.6) where k=2 / is the acoustic wave number, is the acoustic wavelength, a is the root-mean-square (rms) amplitude of the surface waves (2a=crest-to-trough wave height) and is the grazing angle relative to the horizontal plane. When R 1, the sea surface is considered to be acoustically smooth; when R 1, the surface is acoustically rough. Sea-surface wave spectra can be numerically generated by executing available spectral ocean wave models in the hindcast mode. Hindcasting is usually the only means available for obtaining sufficiently long record lengths from which to generate reliable statistics. A statistical analysis of these hindcast data produces probability distributions of critical parameters for use in estimating future sea surface conditions. Eller (1984a) has reviewed the availability of simple surface loss algorithms appropriate for incorporation into propagation models. 3.4.2 Image interference and frequency effects When the surface is smooth, an interference pattern is produced between direct-path sound and sound reflected from the sea surface. The sound reflected from the sea surface may be considered to originate from an image source located on the opposite (mirror image) side of the surface (Fig. 3.8). This image signal will have an amplitude almost equal to that of the incident signal, but will be out of phase. This sound field may be divided into three parts (Fig. 3.9): (1) near field close to the source in which the image source is too far away, and the reflected sound is too weak to produce appreciable interference; (2) interference field in which there are strong peaks and nulls in the signal received as range increases; and (3) far field in which there is an increasingly out-of-phase condition between source and image, and the intensity falls off as the inverse fourth power of the range. This phenomenon, also known as the Lloyd Mirror effect, diminishes with increasing surface roughness. Assuming an acoustically smooth sea surface (R 1) and a shallow source depth (d) in deep water, the ranges r1 and r2 in Fig. 3.9 can be approximated as (Urick, 1982):

where is the acoustic wavelength and H is the receiver depth. The image effect can be used to estimate the depth of a submerged target at short ranges. In Fig. 3.8, if the source is replaced by a submarine which has been ensonified by a single, short pulse, then the depth (d) of the submarine is approximated by (3.7) where c is the speed of sound and t is the difference in time between receipt of the direct and the surfacereflected pulses (Albers, 1965, pp. 50–1). The concept of surface interference can also be used to solve relatively simple propagation problems. The approach is called the ‘method of images’ and is valid for all frequencies. The solution is usually expressed as a sum of the contributions of all images within a multilayered space. This method is usually cumbersome and is commonly employed as a physical model against which to check the results of more elaborate mathematical models (to be discussed in Chapter 4). Kinsler et al. (1982, pp. 427–30) provide a more

50

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.8 Geometry for image interference effect.

Fig. 3.9 Illustration of image interference effects (adapted from Urick, 1979). The transmission anomaly (TA) represents the difference between the observed transmission loss (TL) and losses due to the effects of spherical spreading [TA= 20 log10 (r)—TL], where the range (r) is measured in meters.

detailed discussion of this method together with several examples of its application. Tolstoy and Clay (1966, pp. 33–6) discuss solutions in waveguides. When the sea surface is rough, the vertical motion of the surface modulates the amplitude of the incident wave and superposes its own spectrum as upper and lower sidebands on the spectrum of the incident sound. When there is a surface current, the horizontal motion will appear in the scattered sound and cause a Doppler-shifted and Doppler-smeared spectrum.

BASIC CONCEPTS

51

3.4.3 Bubble layers The presence of bubble layers near the sea surface further complicates the reflection and scattering of sound as a result of the change in sound speed, the resonant characteristics of bubbles and the scattering by the bubbly layers. Hall (1989) developed a comperhensive model of wind-generated bubbles in the ocean. The effects on the transmission of short pulses in the frequency range 1.25–40 kHz were also examined. For long-range propagation, Hall concluded that the decrease in the near-surface sound speed due to bubbles does not significantly affect the intensity of the surface-reflected rays. 3.4.4 Ice interaction Acoustic interaction with an ice canopy is governed by the shape of the under-ice surface and by the compressional wavespeeds (typically 1300– 3900 m/s) and shear wavespeeds (typically 1400–1900 m/s) (see Untersteiner, 1966; Medwin et al., 1988). McCammon and McDaniel (1985) examined the reflectivity of ice due to the absorption of shear and compressional waves. They found that shear wave attenuation is the most important loss mechanism from 20° to 60° incidence for smooth ice at low frequencies ( 2 kHz). In Arctic regions, the presence of a positive-gradient sound speed profile and a rough under-ice surface (with a distribution of large keels) may lead to significant out-of-plane scattering. The acoustic impacts of this scattering are twofold. First, significant beam widening may result from the multiple interactions with the randomly rough surface, similar to that observed in shallow water. Second, the presence of ice keels in the vicinity of the receiver leads to multiple source images or beam-steering errors from the interaction of the acoustic signal with the facets of the local under-ice surface. Because of the overwhelming effect of ice on the propagation of sound in the Arctic, the magnitude of the excess attenuation observed under the ice should be determined by the statistics of the under-ice surface. Ice-ridge models can be categorized according to two classes: discrete models and continuous statistical models. These two classes of models are briefly described below. Discrete ice-ridge models prescribe a representative ridge shape, or an ensemble of ridge shapes, to calculate the statistics of the surface from the discrete statistics of the known ice structure. Continuous statistical ice-ridge models treat the under-ice surface as a stochastic process. This process is then analyzed using the techniques of time-series analysis in which the under-ice surface can be characterized by its autocorrelation function. These models can give a more complete description of the underice roughness than can the discrete models; however, they are limited in application to those surfaces that can be completely specified by a Gaussian depth distribution. The model developed by Diachok (1976) will be discribed since it is considered to be representative of the class of discrete ice-ridge models known to exist and because of its intuitive appeal. The discrete models are also more robust (i.e. require less knowledge of the under-ice surface) than the continuous statistical models. Furthermore, Diachok’s model has been incorporated into existing propagation models with some success. According to Diachok’s model, sea ice may be described as consisting of floating plates, or floes, about 3 m thick, occasionally interrupted by ridges, which are rubble piles formed by collisions and shear interactions between adjacent floes. Ridge dimensions vary widely, but are nominally about 1 m high, 4 m deep and 12 m wide, with the ridge length generally being much greater than the depth or width. A

52

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

Fig. 3.10 Geometrical model of sea-ice ridges (Diachok, 1976).

representative average spacing between ridges (which is random) is about 100 m. Ice-ridge orientation is commonly assumed to be directionally isotropic, although limited empirical data suggest that, at least locally, there may be a preferred orientation. The physical model of reflection developed by Twersky (1957) was used. A comparison between measured contours and simple geometrical shapes suggests that ridge keel contours may reasonably be represented by a half-ellipse (as in Fig. 3.10), and that ridge sail contours may be described using a Gaussian distribution function. The relative dimensions of this geometrical model are indicated in Fig. 3.10. The exact solution of underice scattering off a flat surface with a single semi-elliptical cylindrical boss of infinite extent was developed by Rubenstein and Greene (1991). LePage and Schmidt (1994) extended the applicability of perturbation theory to under-ice scattering at low frequencies (10–100 Hz) by including the scattering of incident acoustic energy into elastic modes, which then propagate through the ice. 3.4.5 Measurements Three basic experimental techniques have been employed to measure forward reflection losses at the sea surface: 1. comparing the amplitude or energy of pulses returned from the surface with that of the direct arrival; 2. using the Lloyd Mirror effect and observing the depth of the minima as the frequency is varied; 3. measuring the attenuation in the surface duct. Based on a compilation of results in the literature by Urick (1982, Chapter 10), it appears that surface losses are less than 1 dB (per bounce) at frequencies below 1 kHz, and rise to about 3 dB (per bounce) at frequencies above 25 kHz. 3.5 SEA-FLOOR BOUNDARY The sea floor affects underwater sound by providing a mechanism for:

BASIC CONCEPTS

1. 2. 3. 4. 5.

53

forward scattering and reflection loss (but complicated by refraction in the bottom) interference and frequency effects attenuation by sediments noise generation at lower frequencies due to seismic activity backscattering and bottom reverberation.

Urick (1982, Chapter 11) provides a comprehensive summary of sound reflection and scattering by the sea floor. The single most important physical property that determines the acoustic characteristics of sediments is their porosity. Items (1) through (3) will be discussed below. Item (4) will be discussed in Chapter 6 and item (5) in Chapter 8. Although the sea floor affects sound in ways similar to that of the sea surface, the return of sound from the bottom is more complex than from the surface for several reasons: (1) the bottom is more variable in composition; (2) the bottom is often stratified (layered) with density and sound speeds (both shear and compressional) varying gradually or abruptly with depth; (3) bottom characteristics (composition and roughness) can vary over relatively short horizontal distances; and (4) sound can propagate through a sedimentary layer and either be reflected back into the water by subbottom layers or be refracted back by the large sound speed gradients in the sediments. These mechanisms can be incorporated into mathematical models through the specification of appropriate ‘boundary conditions’. The complexity of these boundary conditions will depend upon the level of known detail concerning the composition and structure of the sea floor, and also to some degree on the sophistication of the mathematical model being used. The specification of boundary conditions at the bottom has assumed greater importance due to increased interest in the modeling of propagation in shallow-water areas; such propagation, by definition, is characterized by repeated interactions with the bottom boundary. Acoustic interactions with highly variable sea-floor topographies and bottom compositions often necessitate the inclusion of both compressional and shear wave effects, particularly at lower frequencies. A fluid, by definition, cannot support shear stresses; therefore, in modeling acoustic propagation in an ideal (boundless) fluid layer, only compressional wave effects need be considered. As an approximation, saturated sediments are sometimes modeled as a fluid layer in which the sound speed is slightly higher than that of the overlying water column. The basement, however, can support both compressional and shear waves, and rigorous modeling of acoustic waves that interact with and propagate through such media must consider both types of wave effects. As an approximation, shear-wave effects are sometimes included in the form of modified attenuation coefficients. 3.5.1 Forward scattering and reflection loss The ideal forward reflection loss of sound incident on a plane boundary between two fluids characterized only by sound speed and density was originally developed by Rayleigh (1945, Vol. II, p. 78). This model is commonly referred to as Rayleigh’s law. In the simplest model incorporating absorption, the bottom can be taken to be a homogeneous absorptive fluid with a plane interface characterized by its density, sound speed and attenuation coefficient. In the case of sedimentary materials, all three of these parameters are determined by the porosity of the sediments. In underwater acoustics, a common idealized model for the interaction of a point-source field with the sea floor is the so-called Sommerf eld model (after A.N.Sommerfeld). This model consists of an isospeed halfspace water column overlying an isospeed half-space bottom. The bottom has a higher sound speed than the

54

PROPAGATION I. OBSERVATIONS AND PHYSICAL MODELS

water. Thus, a critical angle exists in the plane wave reflection coefficient. For large grazing angles, energy is partially reflected and partially transmitted at the water–bottom interface; for small grazing angles, energy is totally reflected. Energy incident near the critical angle produces a complex phenomenon known as the lateral, or head, wave (Chin–Bing et al., 1982, 1986; Westwood, 1989a). See also the discussion by Clay and Medwin (1977, pp. 262–3). This condition is termed an ‘impedance boundary’. Another commonly assumed boundary condition for the sea floor is the homogeneous Neumann bottom boundary condition. This condition requires that the normal component of the pressure gradient at the bottom be equal to zero. This condition is termed a ‘rigid boundary’. These simplistic models have largely been abandoned in favor of more rigorous geoacoustic models of the sea floor which more properly account for the propagation of sound in sediments (Anderson and Hampton, 1980a, 1980b; Hamilton, 1980). One early attempt at such modeling was provided by Hall and Watson (1967), who developed an empirical bottom reflection loss expression based largely on the results of AMOS (Marsh and Schulkin, 1955). Westwood and Vidmar (1987) have summarized the current state of modeling of acoustic interaction with the sea floor. It is convenient to partition the discussion according to low-frequency and high-frequency bottom interaction. The transition between low and high frequencies is rather imprecise, but can be considered to occur at approximately 200 Hz. At low frequencies and low grazing angles, the acoustic interaction with the sea floor in deep ocean basins is simple and well understood. The relatively long acoustic wavelengths are insensitive to the details of small-scale layering in the sediments. Moreover, for low grazing angles, there is little interaction with the potentially rough substrate interface. Accordingly, the sea floor can be accurately approximated as a horizontally stratified and depth-dependent fluid medium. The major acoustical processes affecting interaction with the sea floor are: (1) reflection and transmission of energy at the water–sediment interface, (2) refraction of energy by the positive sound speed gradient in the sediments, and (3) attenuation within the sediments. Modeling of this interaction is further enhanced by the availability of established methods for estimating the geoacoustic profile (i.e. sound speed, density and attenuation as functions of depth) of deepsea sediments, given the sediment type and physiographic province. In contrast, bottom interaction at high frequencies is not well understood. The relatively short wavelengths are now more sensitive to the small-scale sediment layering. These layers are reported to have an important effect on the magnitude and phase of the plane wave reflection coefficient. Stochastic techniques with which to analyze the effects of the near-surface sediment layering are being developed, but they do not yet incorporate potentially important acoustical processes such as refraction and shear wave generation. Modeling at high frequencies is further frustrated by the high spatial variability of sediment layering. The concept of ‘hidden depths’ (Williams, 1976), which states that the deep ocean sediment structure well below the ray turning point has no acoustical effect, is important because it focuses attention on those lowfrequency processes occurring in the upper regions of the sediments (see Knobles and Vidmar, 1986). Westwood and Vidmar (1987) have developed a ray-theoretical approach, called CAPARAY, for simulating the propagation of broadband signals interacting with a layered ocean bottom. CAPARAY can simulate a time series at a receiver due to an arbitrary source waveform by constructing a frequency domain transfer function from the eigenray characteristics. Geoacoustic models have been developed to account for bottom interaction processes in propagation models. As summarized by Holland and Brunson (1988), geoacoustic models of marine sediments can be formulated in one of three ways: (1) by empirically relating geoacoustic and geophysical properties of the sediments (e.g. Hamilton, 1980); (2) by using the Biot–Stoll model to relate sediment geoacoustic

BASIC CONCEPTS

55

Fig. 3.11 Thin layer model for sediment reflected and refracted paths (McCammon, 1988; J. Geophys. Res., 93, 2363– 9; published by the American Geophysical Union).

properties to geophysical properties on the basis of physical principles; and (3) by using an inversion technique to generate sediment geophysical parameters from bottom loss measurements [e.g. McCammon, 1991; Hovem et al., 1991 (see especially Section 3, Modelling and inversion techniques); Rajan, 1992; Dosso et al., 1993; Hovem, 1993; Frisk, 1994]. The Biot–Stoll model (Biot, 1956a, b; Stoll, 1974, 1980, 1989) provides a comprehensive description of the acoustic response of linear, porous materials containing a compressible pore fluid. The model predicts two types of compressional waves and one shear wave. Recent applications in underwater acoustics and references to the key historical literature are provided by Beebe et al. (1982) and Holland and Brunson (1988). Routine operational employment of this model is hampered by the input of more than a dozen geophysical parameters, some of which are difficult to obtain even in laboratory environments. McCammon (1988) described the development of a geoacoustic approach to bottom interaction called the thin layer model. This model, which is based on an inversion technique, contains a thin surificial layer, a fluid sediment layer and a reflecting subbottom half-space. There are ten input parameters: sediment density, thickness, sound speed gradient and curvature, attenuation and attenuation gradient; thin layer density and thickness; basement reflectivity; and water–sediment velocity ratio (Fig. 3.11). The model generates bottom loss curves as a function of grazing angle over the frequency range 50–1500 Hz. The model makes several assumptions: it relies upon the ‘hidden depths’ concept of Williams (1976); the sediments are isotropic; the roughness of the sediment and basement interfaces, as well as multiple scattering within the layers, is neglected; and shear wave propagation is ignored. Sample outputs from this thin layer model are presented in Fig. 3.12. A ratio (cs/cw)>1 (where cs is the sound speed in the upper sediment and cw is the sound speed at the base of the water column) predicts a critical angle [ =cos−1(cw/cs)] below which most of the incident energy is reflected; that is, the bottom loss is nearly zero. By comparison, a ratio (cs/cw)